Loading [MathJax]/jax/element/mml/optable/GeneralPunctuation.js
Research article Special Issues

Some evaluations of the fractional p-Laplace operator on radial functions

  • We face a rigidity problem for the fractional p-Laplace operator to extend to this new framework some tools useful for the linear case. It is known that (Δ)s(1|x|2)s+ and Δp(1|x|pp1) are constant functions in (1,1) for fixed p and s. We evaluated (Δp)s(1|x|pp1)s+ proving that it is not constant in (1,1) for some p(1,+) and s(0,1). This conclusion is obtained numerically thanks to the use of very accurate Gaussian numerical quadrature formulas.

    Citation: Francesca Colasuonno, Fausto Ferrari, Paola Gervasio, Alfio Quarteroni. Some evaluations of the fractional p-Laplace operator on radial functions[J]. Mathematics in Engineering, 2023, 5(1): 1-23. doi: 10.3934/mine.2023015

    Related Papers:

    [1] Nicola Abatangelo, Sven Jarohs, Alberto Saldaña . Fractional Laplacians on ellipsoids. Mathematics in Engineering, 2021, 3(5): 1-34. doi: 10.3934/mine.2021038
    [2] Federico Cluni, Vittorio Gusella, Dimitri Mugnai, Edoardo Proietti Lippi, Patrizia Pucci . A mixed operator approach to peridynamics. Mathematics in Engineering, 2023, 5(5): 1-22. doi: 10.3934/mine.2023082
    [3] María Ángeles García-Ferrero, Angkana Rüland . Strong unique continuation for the higher order fractional Laplacian. Mathematics in Engineering, 2019, 1(4): 715-774. doi: 10.3934/mine.2019.4.715
    [4] Antonio Iannizzotto, Giovanni Porru . Optimization problems in rearrangement classes for fractional $ p $-Laplacian equations. Mathematics in Engineering, 2025, 7(1): 13-34. doi: 10.3934/mine.2025002
    [5] Antonio Greco, Francesco Pisanu . Improvements on overdetermined problems associated to the $ p $-Laplacian. Mathematics in Engineering, 2022, 4(3): 1-14. doi: 10.3934/mine.2022017
    [6] Chao Xia, Jiabin Yin . Two overdetermined problems for anisotropic $ p $-Laplacian. Mathematics in Engineering, 2022, 4(2): 1-18. doi: 10.3934/mine.2022015
    [7] María Medina, Pablo Ochoa . Equivalence of solutions for non-homogeneous $ p(x) $-Laplace equations. Mathematics in Engineering, 2023, 5(2): 1-19. doi: 10.3934/mine.2023044
    [8] Catharine W. K. Lo, José Francisco Rodrigues . On the obstacle problem in fractional generalised Orlicz spaces. Mathematics in Engineering, 2024, 6(5): 676-704. doi: 10.3934/mine.2024026
    [9] Patrizia Pucci, Letizia Temperini . On the concentration–compactness principle for Folland–Stein spaces and for fractional horizontal Sobolev spaces. Mathematics in Engineering, 2023, 5(1): 1-21. doi: 10.3934/mine.2023007
    [10] Peter Bella, Mathias Schäffner . Local boundedness for $ p $-Laplacian with degenerate coefficients. Mathematics in Engineering, 2023, 5(5): 1-20. doi: 10.3934/mine.2023081
  • We face a rigidity problem for the fractional p-Laplace operator to extend to this new framework some tools useful for the linear case. It is known that (Δ)s(1|x|2)s+ and Δp(1|x|pp1) are constant functions in (1,1) for fixed p and s. We evaluated (Δp)s(1|x|pp1)s+ proving that it is not constant in (1,1) for some p(1,+) and s(0,1). This conclusion is obtained numerically thanks to the use of very accurate Gaussian numerical quadrature formulas.



    In this paper we wish to investigate, in a nonlocal nonlinear framework, some tools that have proved to be particularly useful for obtaining symmetry results for local operators.

    It is well known that one of the crucial steps for applying the moving plane method to overdetermined problems à la Serrin is via a comparison principle. In the nonlocal setting there are, in the literature, several versions of comparison principles: in the linear case p=2, they follow by linearity from the maximum principle, while in the nonlinear case p2, they are more difficult to obtain. Strong maximum principles for fractional Laplacian-type operators have been proved in [19], a weak maximum principle for antisymmetric solutions of problems governed by the fractional Laplacian can be found in [10] (see also [16] for more general nonlocal operators), and a version of the strong maximum principle in the case of nonlocal Neumann boundary conditions can be found in [5]. For the fractional p-Laplacian operator, we refer to [17] for a weak comparison principle (see also [13]), and to [12,15] for two different versions of the strong comparison principle; while some versions of the strong maximum principle and Hopf lemma can be found in [4,7]. In the first part of this paper we revisit some results concerning the comparison principle for the fractional p-Laplace operator in bounded domains and prove a slightly new version of the strong comparison principle in Theorem 2.1.

    In the second part of the paper we address the study of the p-fractional torsion problem

    {(Δp)su=1in B,u=0in RNB, (1.1)

    where s(0,1), p>1, BRN (N1) is a ball,

    (Δp)su(x):=cN,s,plimε0+(Bε(x))c|u(x)u(y)|p2(u(x)u(y))|xy|N+psdy,xB (1.2)

    denotes the fractional p-Laplace operator, and cN,s,p=sp2(1s)22s1πN12Γ(N+sp2)Γ(p+12)Γ(2s)>0 is a normalization constant whose exact value plays a role in the limit cases s1 and p2, cf. [8,Lemma 5.1]. Such a problem admits a unique solution, which is radial and radially non-increasing, cf. [13,Lemma 4.1], but whose analytic expression is not known in the nonlocal nonlinear case: s(0,1) and p2.

    On the other hand, in the local case s=1, for the p-Laplace operator Δpu=div(|u|p2u), it is easy to prove that the function (1|x|m), with m=pp1, has constant p-Laplacian in (1,1), see for instance [6]. Moreover, in the linear case p=2, it has been proved that (1x2)s+ satisfies (Δ)s(1x2)s+=Const. in (1,1), see [9]. In view of these two results, and recalling that (Δp)su(x)Δpu(x), when s1, see [8,14], as well as, of course, that (Δ2)su(x)=(Δ)su(x), it would be interesting to check whether the function (1|x|m)s+ may satisfy the equation

    (Δp)s(1|x|m)s+=Const.>0,

    for every x(1,1)R. In fact, the construction of the solution of the problem (1.1) would follow easily by a homogeneity argument. This result however does not hold true. As a matter of fact, we prove that there exist p>2, s(0,1), x1,x2(1,1) such that x1x2 and (Δp)s(1|x1|m)s+(Δp)s(1|x2|m)s+. Our proof follows by investigating the value of

    ((Δp)s(1|x|m)s+)|x=0=2c1,s,p(1sp+10(1(1ym)s)p1y1+spdy),

    where the value of c1,s,p is given below formula (1.2), for N=1.

    The paper is organized as follows. In Section 2 we deduce a strong comparison principle that holds for the fractional p-Laplace operator in any dimension N1. Notice that, in the local case, a similar result has been proved only in dimension N=2, see [18]. In Section 3 we prove, by following a different strategy with respect to [9], that the s-fractional Laplace operator of (1x2)s+ in (1,1) is constant. In Section 4 we prepare the ground for a numerical evaluation of the s-fractional p-Laplacian of (1|x|m)s+, proving integrability properties, see Propositions 4.1 and 4.3, that are useful to yield error estimates for our numerical integration formaulae. Finally, in Section 5 we show, by computing numerically the integral in (1.2), that there exist p2 and s(0,1) such that the s-fractional p-Laplace operator of (1|x|m)s+ is not constant in (1,1).

    The software written to produce the numerical results of the present paper is freely available on github at the URL https://github.com/pgerva/fractional-p-laplace.git.

    In this section, we consider the following system of inequalities

    {(Δp)su+q(x)|u|p2u(Δp)sv+q(x)|v|p2vin Ω,uvin RN, (2.1)

    where s(0,1), p>1, ΩRN (N1) is a bounded domain, qL(Ω), and (Δp)s denotes the fractional p-Laplacian, which, on smooth functions u, can be written as

    (Δp)su(x):=cN,s,plimε0+(Bε(x))c|u(x)u(y)|p2(u(x)u(y))|xy|N+psdy,xΩ, (2.2)

    where cN,s,p>0 is the usual normalization constant introduced in the Introduction.

    We will prove the following strong comparison principle.

    Theorem 2.1. Let (u,v) be a weak solution of (2.1). If u,vC(Ω) and

    ΩΩ||v(x)v(y)|p2(v(x)v(y))|u(x)u(y)|p2(u(x)u(y))||xy|N+spdxdy<, (2.3)

    then either u<v in Ω or uv in RN.

    The proof of this theorem is based on an argument first introduced in [18]. We observe that the previous strong comparison principles for (Δp)s, [15,Theorem 1.1] and [12,Theorem 2.7], require different regularity assumptions on u and v and use a different proof technique. Before proving Theorem 2.1, we introduce the functional spaces and the main definitions that will be useful to work with weak solutions, and prove a preliminary lemma.

    For every s(0,1) and p(1,), we define

    Ws,p(RN):={uLp(RN):RNRN|u(x)u(y)|p|xy|N+spdxdy<},
    Ws,p0(Ω):={uWs,p(RN):u0 in RNΩ},
    ˜Ws,p(Ω):={uLploc(RN):ΩRN|u(x)u(y)|p|xy|N+spdxdy<}.

    Definition 2.2. A function u˜Ws,p(Ω) is a weak solution of (Δp)su()0 in Ω if

    RNRN|u(x)u(y)|p2(u(x)u(y))(φ(x)φ(y))|xy|N+spdxdy()0

    holds for every 0φWs,p0(Ω). Consequently, a function u˜Ws,p(Ω) is a weak solution of (Δp)su=0 in Ω if

    RNRN|u(x)u(y)|p2(u(x)u(y))(φ(x)φ(y))|xy|N+spdxdy=0

    holds for every φWs,p0(Ω).

    A couple (u,v)(˜Ws,p(Ω))2 is a weak solution of (2.1) if the inequality

    cN,s,pRNRN|u(x)u(y)|p2(u(x)u(y))(φ(x)φ(y))|xy|N+spdxdy+Ωq(x)|u(x)|p2u(x)φ(x)dxcN,s,pRNRN|v(x)v(y)|p2(v(x)v(y))(φ(x)φ(y))|xy|N+spdxdy+Ωq(x)|v(x)|p2v(x)φ(x)dx

    holds for every 0φWs,p0(Ω), and uv a.e. in RN.

    Lemma 2.3. If fL1loc(Ω) is such that

    Ωf(x)φ(x)dx0forevery0φCc(Ω),

    then f0 a.e. in Ω.

    Proof. Let Ω=n=1Ωn, with ΩnΩn+1, and let Kn be compact sets such that

    KnΩnKn+1for every nN. (2.4)

    Since fL1loc(Ω), fL1(Kn) and consequently, by (2.4), fL1(Ωn) for every n. In particular, f+,fL1(Ωn) and also 1{f>0}L1(Ωn). Now, fix nN. By density, there exists a sequence (φj)Cc(Ωn) such that φj1{f>0} in L1(Ωn). Therefore, passing if necessary to a subsequence, and using the Dominated Convergence Theorem, we get

    limjΩnfφjdx=Ωnlimj(f+f)φjdx=Ωn{f>0}fdx. (2.5)

    Now, by assumption, for every jN, Ωnfφjdx=Ωfφjdx0, being φjCc(Ωn)Cc(Ω). Hence, by (2.5),

    Ωn{f>0}fdx0. (2.6)

    We can now pass to the limit in n to obtain

    0limnΩn{f>0}fdx=limn{f>0}f1Ωndx={f>0}f1Ωdx={f>0}fdx0,

    where we have used (2.6), 1Ωn1Ω a.e. and the Monotone Convergence Theorem. This immediately gives |{f>0}|=0 and concludes the proof.

    Proof of Theorem 2.1. We introduce the notation A(h):=|h|p2h for every hR. Since (u,v) is a weak solution of (2.1), we get for every 0φCc(Ω)

    I:=1cN,s,pΩq(x)(|u(x)|p2u(x)|v(x)|p2v(x)|)φ(x)dxRNRN{|v(x)v(y)|p2(v(x)v(y))|u(x)u(y)|p2(u(x)u(y))}φ(x)φ(y)|xy|N+spdxdy=RNRN{A(v(x)v(y))A(u(x)u(y))}φ(x)φ(y)|xy|N+spdxdy. (2.7)

    On the other hand, let ut:=tv+(1t)u for every t[0,1] and w:=vu, then by straightforward calculations we have

    A(v(x)v(y))A(u(x)u(y))=10ddtA(ut(x)ut(y))dt=(p1)(10|ut(x)ut(y)|p2dt)(w(x)w(y))=:a(x,y)(w(x)w(y)).

    We observe that a(x,y)=a(y,x) for every x,yRN. Hence, continuing the estimate in (2.7) and using that φ0 in RNΩ, we have

    IRNRNa(x,y)(w(x)w(y))(φ(x)φ(y))|xy|N+spdxdy=RNΩΩdxdy+ΩΩdxdy+ΩRNΩdxdy.

    By the symmetry of a(,), we notice that the first and the third integral in the last expression are equal, so that we can write

    I2RNΩ(Ωa(x,y)(w(x)w(y))φ(x)|xy|N+spdx)dy+ΩΩa(x,y)(w(x)w(y))(φ(x)φ(y))|xy|N+spdxdy=RN(Ωa(x,y)(w(x)w(y))φ(x)|xy|N+spdx)dy+RNΩ(Ωa(x,y)(w(x)w(y))φ(x)|xy|N+spdx)dyΩ(Ωa(x,y)(w(x)w(y))φ(y)|xy|N+spdx)dy. (2.8)

    As for the last integral, in view of (2.3), we can manipulate it in the following way

    Ω(Ωa(x,y)(w(x)w(y))φ(y)|xy|N+spdx)dy=Ω(Ωa(x,y)(w(x)w(y))φ(y)|xy|N+spdy)dx=Ω(Ωa(x,y)(w(x)w(y))φ(x)|xy|N+spdx)dy,

    therefore, we can sum up the last two integrals in (2.8) to get in conclusion

    I2RN(Ωa(x,y)(w(x)w(y))|xy|N+spφ(x)dx)dy=2Ω(RNa(x,y)(w(x)w(y))|xy|N+spdy)φ(x)dx (2.9)

    for every 0φCc(Ω). Arguing in a similar way, we can re-write the integral I as follows

    I=1cN,s,pΩq(x)(A(u(x))A(v(x)))φ(x)dx=1cN,s,pΩq(x)(10ddtA(ut(x))dt)φ(x)dx=Ωq(x)((p1)10|ut(x)|p2dtcN,s,p)w(x)φ(x)dx=:Ωq(x)b(x)w(x)φ(x)dx. (2.10)

    Combining together (2.9) and (2.10), we get

    Ω(2RNa(x,y)(w(x)w(y))|xy|N+spdy+q(x)b(x)w(x))φ(x)dx0

    for every 0φCc(Ω). Thus, by Lemma 2.3, this implies that

    2RNa(x,y)w(x)w(y)|xy|N+spdyq(x)b(x)w(x)for a.e. xΩ.

    Now, suppose by contradiction that there exists x0Ω such that w(x0)=0, then

    RNa(x0,y)w(y)|x0y|N+spdy0. (2.11)

    Since w=vu0 a.e. in RN and a(x,y)0 for a.e. x,yRN, (2.11) implies that a(x0,y)w(y)=0 for a.e. yRN.

    We are now ready to conclude. We observe that, if a(x0,y)=0 for some yRN, then w(y)=0. Indeed, by straightforward calculations, if

    a(x0,y)=(p1)10|u(x0)u(y)+t(u(y)v(y))|p2dt=0,

    then

    u(x0)u(y)+t(u(y)v(y))=0for every t[0,1],

    which gives u(y)=v(y), or equivalently w(y)=0. So, we have proved that a(x0,y)w(y)=0 for a.e. yRN is equivalent to w(y)=0 for a.e. yRN, which concludes the proof.

    Arguing as in [17,Lemma 9], we have the following weak comparison principle. We stress that, with respect to Theorem 2.1, we need to ask u to be continuous in the whole space and q to be non-negative.

    Lemma 2.4. Let (u,v) be a weak solution of

    {(Δp)su+q(x)|u|p2u(Δp)sv+q(x)|v|p2vinΩ,uvinRNΩ,

    where 0qL(Ω).If u,vC(RN), then uv also in Ω.

    Proof. Reasoning as in the first part of the proof of Theorem 2.1, we get for every 0φWs,p0(Ω)

    RNRNa(x,y)(w(x)w(y))φ(x)φ(y)|xy|N+spdxdyΩq(x)b(x)w(x)φ(x)dx, (2.12)

    with the same definitions for a, b, and w=vu. Now, following the idea in [17,Lemma 9], we choose φ:=(uv)+=w and observe that

    wφ=(w+w)w=(w)20.

    Hence, putting φ=w, and using that q0, we get from (2.12)

    RNRNa(x,y)(w(x)w(y))w(x)w(y)|xy|N+spdxdy0.

    The proof now can be completed exactly as in [17,Lemma 9].

    Combining the previous lemma with Theorem 2.1, we get the following.

    Corollary 2.5. Let (u,v) be a weak solution of

    {(Δp)su+q(x)|u|p2u(Δp)sv+q(x)|v|p2vinΩ,uvinRNΩ,

    where 0qL(Ω). If u,vC(RN) and (2.3) holds, then either u<v in Ω or uv in RN.

    Let s(0,1). In this section we consider the following problem

    {(Δ)su=1in BRu=0in RNBR, (3.1)

    where BRRN is the ball of radius R centered at the origin.

    Remark 3.1. We observe that, at least formally, the N-dimensional fractional Laplacian (Δ)sN of a function u:RNR can be expressed in terms of the 1-dimensional fractional Laplacian (Δ)s1 of related functions of one variable. Indeed, denoting simply cN,s:=cN,s,2, we get for every xRN

    1cN,s(Δ)sNu(x)=limε0+RNBε(0)u(x)u(y)|xy|N+2sdy=limε0++ε(Bt(x)u(x)u(y)|xy|N+2sdσ(y))dt=limε0++ε(B1(x)u(x)u(xtν)t2s+1dσ(ν))dt=limε0+{B1(x){xN>0}(+εu(x)u(xtν)t2s+1dt)dσ(ν)=+B1(x){xN<0}(+εu(x)u(xtν)t2s+1dt)dσ(ν)}=limε0+B1(x){xN>0}(R(ε,ε)u(x)u(xtν)|t|2s+1dt)dσ(ν)=B1(x){xN>0}(limε0+R(ε,ε)u(x)u(xtν)|t|2s+1dt)dσ(ν)=B1(x){xN>0}1c1,s(Δ)s1ψν,x(0)dσ(ν),

    where ψν,x(t):=u(xtν) for every tR. In particular, for uN(x)=(1|x|2)s+, ψν,x(t)=u1(|xtν|), and so, once it is proved that (Δ)s1u1 is constant, one has immediately that also (Δ)sNuN is constant.

    In the light of the previous remark, from now on in the paper we consider only the case of dimension N=1, and drop all subscripts referring to the dimension. Moreover, for the sake of simplicity, we take the radius R to be 1. In this setting, we give an alternative proof of the fact that the solution of (3.1) is given by vs(x):=sin(πs)πcs(1x2)s+, where cs is the normalization constant for the fractional Laplacian in dimension one and is given by cs:=22sπΓ(1+2s2)Γ(1s)s, cf. for instance [1,Remark 3.11]. We refer to [9] for a previous proof.

    Theorem 3.2. Let N=1 and vs(x):=sin(πs)πcsus(x), with us(x):=(1x2)s+. Then vs is a Cs([1,1]) solution of (3.1).

    Proof. For every xR(1,1), us(x)=0. Moreover, for every x(1,1),

    (Δ)sus(x)=cslimε0+R(xε,x+ε)us(x)us(y)|xy|1+2sdy=cslimε0+R(xε,x+ε)us(x)us(y)|x+y|1+2sdy=cslimε0+R(xε,x+ε)us(x)us(z)|xz|1+2sdz=(Δ)sus(x).

    Now, let x(0,1) and ε>0, then

    R(xε,x+ε)us(x)us(y)|xy|1+2sdy=xε1(1x2)s(1y2)s|xy|1+2sdy+1x+ε(1x2)s(1y2)s|xy|1+2sdy+(1x2)sR(1,1)1|xy|1+2sdy=:I1(x)+I2(x)+I3(x).

    As for the last integral, we immediately get

    I3(x)(1x2)s=1x1+2s(11|1y/x|1+2sdy+11|1y/x|1+2sdy)=12sx2s[(1+1x)2s+(1x1)2s]=12s[1(1x)2s+1(1+x)2s].

    We manipulate and integrate by parts I1(x) to obtain

    I1(x)=(1x2)sxε11(xy)1+2sdyxε1(1y2)s(xy)1+2sdy=(1x2)s2s(1ε2s1(1+x)2s){(1(xε)2)s2sε2s+xε1(1y2)s1(xy)2sydy}. (3.2)

    Similarly, for I2(x) we have

    I2(x)=(1x2)s1x+ε1(yx)1+2sdy1x+ε(1y2)s(yx)1+2sdy=(1x2)s2s(1ε2s1(1x)2s){(1(x+ε)2)s2sε2s1x+ε(1y2)s1(yx)2sydy}. (3.3)

    Now, it is straightforward to see that, as ε0+,

    (1(xε)2)s=(1x2)s+2sx(1x2)1sε+O(ε2),(1(x+ε)2)s=(1x2)s2sx(1x2)1sε+O(ε2).

    Thus, combining them with (3.2) and (3.3), we obtain as ε0+

    I1(x)=(1x2)s2s(1+x)2sε12sx(1x2)1sxε1(1y2)s1(xy)2sydy+O(ε2(1s))

    and

    I2(x)=(1x2)s2s(1x)2s+ε12sx(1x2)1s+1x+ε(1y2)s1(yx)2sydy+O(ε2(1s)).

    Altogether, we have

    (Δ)sus(x)=cslimε0+(I1(x)+I2(x)+I3(x))=cslimε0+(xε1(1y2)s1(xy)2sydy+1x+ε(1y2)s1(yx)2sydy+O(ε2(1s))). (3.4)

    Now, we distinguish two cases depending on whether s(0,12) or s[12,1).

    Case s(0,12)_. In this case, all integrals involved in the fractional Laplacian of us are convergent. So, in this case, we have

    (Δ)sus(x)=cs(x1(1y2)s1(xy)2sydy+1x(1y2)s1(yx)2sydy).

    Now, by the following change of variable t=xy1xy, we have

    x1(1y2)s1(xy)2sydy=1(1x2)2s110(xt)((1tx)2(xt)2)s1t2s(1tx)dt=1(1x2)2s110(tx)[(1t2)(1x2)]s1t2s(1tx)dt=1(1x2)s10(tx)(1t2)s1t2s(1tx)dt

    and similarly

    1x(1y2)s1(yx)2sydy=1(1x2)s01(xt)(1t2)s1t2s(1tx)dt=1(1x2)s10(x+t)(1t2)s1t2s(1+tx)dt.

    So that, summing up, we have

    (Δ)sus(x)=cs(1x2)s10(1t2)s1t2s[tx1tx+t+x1+tx]dt=2cs(1x2)1s10(1t2)s1t2s1(1t2x2)dt.

    We can now integrate using power series to get

    (Δ)sus(x)=2cs(1x2)1s10(1t2)s1t2s1k=0(tx)2kdt=2cs(1x2)1sk=0(x2k10(1t2)s1t2s12kdt)=2cs(1x2)1sk=0(x2kΓ(s)Γ(ks+1)2Γ(k+1))=csΓ(s)Γ(1s)(1x2)1sk=0(1)k(s1k)x2k=csΓ(s)Γ(1s)(1x2)1s(1x2)s1=csΓ(s)Γ(1s)=csπsinsπ

    where we have calculated the integral 10(1t2)s1t2s12kdt using the change of variables τ=t2, dt=12τdτ, the definition of the Beta function, and the relation between the Beta and the Gamma functions B(x,y)=Γ(x)Γ(y)Γ(x+y), and we have used that Γ(k+1)=k!, the following property of the Gamma function (cf. for instance [20,formula (1.47)]), with z=1s:

    Γ(z+k)Γ(z)=(z)kfor z>k,z0,1,2,,

    the relation between definition of the Pochhammer symbol and the binomial coefficient (cf. for instance [20,formula (1.48)])

    (zk)=(1)k(z)kk!,

    and finally that Γ(s)Γ(1s)=πsinsπ. The conclusion, in this case, follows at once for vs, using the linearity of the fractional Laplacian.

    Cases[12,1)_. In this case, the situation is technically more involved because, when considered individually, the integrals that appear in (3.4) are not convergent and one has to take carefully into account the cancellations. Using the change of variables t:=xy1xy, we get

    xε1(1y2)s1(xy)2sydy=1(1x2)s(1ε1x(xε)(1t2)s1t2s1(1tx)dtx1ε1x(xε)(1t2)s1t2s(1tx)dt),

    where the first integral on the right-hand side is convergent as ε0+. Similarly,

    1x+ε(1y2)s1(yx)2sydy=1(1x2)s(1ε1x(x+ε)(1t2)s1t2s1(1+tx)dt+x1ε1x(x+ε)(1t2)s1t2s(1+tx)dt),

    where, again, the first integral on the right-hand side is convergent as ε0+. Therefore, (3.4) can be re-written in the form

    (Δ)sus(x)=cs(1x2)s10(1t2)s1t2s1(11tx+11+tx)dt+csx(1x2)slimε0+{1ε1x(x+ε)(1t2)s1t2s(1+tx)dt1ε1x(xε)(1t2)s1t2s(1tx)dt+O(ε2(1s))}=:J1(x)+J2(x).

    As for J1(x) one can integrate using power series as already done in the case s<1/2 and obtain

    J1(x)=2cs(1+x2)s10(1t2)s1t2s1k=0(tx)2kdt=2cs(1+x2)sk=0(x2k10(1t2)s1t2s12kdt)=csΓ(s)Γ(1s)1x2.

    We now consider J2(x). We use again power series to get for every ε>0

    1ε1x(x+ε)(1t2)s1t2s11+txdt=k=0(1)kxk1ε1x(x+ε)(1t2)s1t2skdt

    and similarly

    1ε1x(xε)(1t2)s1t2s11txdt=k=0xk1ε1x(xε)(1t2)s1t2skdt.

    We observe that the integrals in the series are all convergent as ε0+ except for the first ones, where k=0. So, we isolate these first terms and calculate, for every ε>0:

    1ε1x(x+ε)(1t2)s1t2s(1+tx)dt1ε1x(xε)(1t2)s1t2s(1tx)dt=1ε1x(x+ε)(1t2)s1t2sdt1ε1x(xε)(1t2)s1t2sdt+k=1{(1)k1ε1x(x+ε)(1t2)s1t2skdt1ε1x(xε)(1t2)s1t2skdt}xk

    Now, let F(t) be a primitive of f(t):=(1t2)s1t2s, then clearly

    1ε1x(x+ε)f(t)dt1ε1x(xε)f(t)dt=F(ε1x(xε))F(ε1x(x+ε)). (3.5)

    Such a primitive can be expressed in terms of the hypergeometric function 2F1 as follows:

    F(t)=t12s2F1(12s,1s,32s;t2)12s=t12s12s(1+O(t2))as t0.

    Inserting this expansion in (3.5), by straightforward calculations we get

    1ε1x(x+ε)f(t)dt1ε1x(xε)f(t)dt=2x(ε1x2)22s+o(ε22s)=O(ε2(1s)).

    In particular, limε0+(1ε1x(x+ε)f(t)dt1ε1x(xε)f(t)dt) is finite. Moreover, we show below that it is finite also the sum of the following series

    k=1limε0+{(1)k1ε1x(x+ε)(1t2)s1t2skdt1ε1x(xε)(1t2)s1t2skdt}xk=k=1((1)k1)xk10(1t2)s1t2skdt=Γ(s)2k=0((1)k+11)xk+1Γ(k2s+22)Γ(k+22)=Γ(s)2xk=0(2)x2kΓ(2k2s+22)Γ(2k+22)=Γ(s)xk=0x2kΓ(ks+1)Γ(k+1)=Γ(s)Γ(1s)xk=0(s1k)(1)kx2k,

    where we have calculated the integral 10(1t2)s1t2skdt using the change of variables τ=t2, dt=12τdτ, the definition of the Beta function, and the relation between the Beta and the Gamma functions, and we have used the sum of the series k=0x2kΓ(ks+1)Γ(k+1) already calculated for the case s<1/2. Therefore, it is possible to pass to the limit as ε0+ in the expression of J2(x) under the series, to get altogether,

    J2(x)=csx(1x2)slimε0+{1ε1x(x+ε)(1t2)s1t2s(1+tx)dt1ε1x(xε)(1t2)s1t2s(1tx)dt+O(ε2(1s))}=csx(1x2)s{limε0+(O(ε2(1s)))+k=1((1)k1)xk10(1t2)s1t2skdt}=csΓ(s)Γ(1s)x2(1x2)sk=0(s1k)(1)kx2k=csΓ(s)Γ(1s)x2(1x2)s(1x2)s1=csΓ(s)Γ(1s)x2(1x2).

    In conclusion,

    (Δ)su(x)=J1(x)+J2(x)=csΓ(s)Γ(1s)1x2csΓ(s)Γ(1s)x2(1x2)=csΓ(s)Γ(1s),

    which proves the thesis also in this case.

    Let s(0,1), p>1, and denote by

    us,p(x):=(1|x|m)s+,m:=pp1.

    Having in mind that, for p=2, the fractional Laplacian of us,2(x)=(1|x|2)s+ is constant in (1,1), see for instance Section 3, and that Δp(1|x|pp1) is constant in (1,1), see for instance [6], it is tempting to conjecture that also (Δp)sus,p is constant in (1,1). In the next section we verify numerically that this conjecture is false.

    To this aim, we first prove in this section some preliminary results.

    Proposition 4.1. For every s(0,1) and p>1, us,pWs,p(R).

    Proof. Clearly, us,p(x)=(1|x|pp1)s+Lp(R). To prove that us,pWs,p(R), we need to show that I:=RR|us,p(x)us,p(y)|p|xy|1+spdxdy<. We write the integral under consideration as follows:

    I=2R(1,1)(11(1|y|m)sp|xy|1+spdy)dx+11(11|(1|x|m)s(1|y|m)s|p|xy|1+spdy)dx=:2I1+I2.

    The integral I1 is convergent. Indeed, arguing as for the integral I3(x) in the proof of Theorem 3.2, we get

    I1=1sp11(1(1+y)sp+1(1y)sp)(1|y|m)spdy.

    Moreover, (1ym)sp(1y)spmsp as y1, and similarly, (1|y|m)sp(1+y)sp is bounded in a neighborhood of y=1. To study the convergence of the integral I2, it is more convenient to change variable and put t=xy1xy in the inner integral, to get

    I2=11(11|(1|x|m)s(1|xt1tx|m)s|p1|t|1+sp(1tx)1spdt)1(1x2)spdx.

    Now, as t1, the integrand of the inner integral has the following asymptotics

    |(1|x|m)s(1|xt1tx|m)s|p|t|1+sp(1tx)1sp(1|x|m)sp(1x)1sp

    and so, for t(1ε,1), I2 has the same behavior of

    11(1|x|m)sp(1x)1sp(1x2)spdx,

    which, in view of the fact that

    1|x|m={m(x+1)+o(x+1)as x1,m(1x)+o(x1)as x1, (4.1)

    is convergent.

    On the other side, as t0,

    xt1tx=(xt)(1+tx+o(t))=xt(1x2)+o(t)|xt1tx|m=|xt(1x2)+o(t)|m=|x|m(1m1x2xt+o(t))(1|xt1tx|m)s=(1|x|m)s(1+ms|x|m(1x2)x(1|x|m)t+o(t))|(1|x|m)s(1|xt1tx|m)s|α(1|x|m)sα(ms)α|x|(m1)α(1x2)α(1|x|m)α|t|α, (4.2)

    for any α>0. Therefore, the integrand of the inner integral (in dt) of I2 has the following asymptotics as t0

    |(1|x|m)s(1|xt1tx|m)s|p|t|1+sp(1tx)1sp(1|x|m)sp(ms)p|x|(m1)p(1x2)p(1|x|m)p1|t|1p(1s),

    and so the integral in dt is convergent. Finally, for t in a neighborhood of 0, I2 has the same behavior of

    11|x|(m1)p(1x2)p(1s)(1|x|m)p(1s)dx,

    which again in view of (4.1) converges.

    Remark 4.2. Arguing as in the first part of the proof of Theorem 3.2, for the fractional p-Laplacian of us,p=(1|x|m)s+, it is possible to calculate explicitly its value at x=0. Indeed, denoting by cs,p the normalization constant involved in the definition of the fractional p-Laplacian in dimension 1, we get for every x(1,1)

    (Δp)sus,p(x)cs,p=|us,p(x)|p2us,p(x)R(1,1)1|xy|1+spdy+limε0+(1,1)(ε,ε)|us,p(x)us,p(y)|p2(us,p(x)us,p(y))|xy|1+spdy=us,p(x)p1(11|xy|1+spdy+11|xy|1+spdy)+limε0+(1,1)(ε,ε)|us,p(x)us,p(y)|p2(us,p(x)us,p(y))|xy|1+spdy=(1|x|m)s(p1)sp(1(1+x)sp+1(1x)sp)+limε0+(1,1)(ε,ε)|us,p(x)us,p(y)|p2(us,p(x)us,p(y))|xy|1+spdy.

    At x=0, the previous expression becomes

    (Δp)sus,p(0)cs,p=2sp+limε0+(1,1)(ε,ε)(1(1|y|m)s)p1|y|1+spdy=2sp+210(1(1ym)s)p1y1+spdy, (4.3)

    where the integral on the last line is meant in the generalized sense, it is convergent and, at least for some values of s and p, can be explicitly expressed in terms of special functions. The value in (4.3) can be taken as reference value for the numerical analysis.

    Proposition 4.3. For every p>1, s(0,11/p), and for every x(1,1), the function

    gx(y):=|us,p(x)us,p(y)|p2(us,p(x)us,p(y))|xy|1+sp (4.4)

    has finite integral over (1,1). Moreover, if 2p(1s)<0, gx belongs to the space Wr,q(R) whenever q1 and 0r<min{1,p(1s)2}.

    Proof. Fix any x(1,1). Via the usual change of variable t=xy1xy, we get

    11gx(y)dy=11|(1|x|m)s(1|xt1tx|m)s|p2[(1|x|m)s(1|xt1tx|m)s]|t|1+sp(1x21tx)spdt=:11fx(t)dt.

    Using (4.2), with α=p2, we have that fx(t)c(x)|t|p2t|t|1+sp as t0, and so the integral is finite. Now, in order to prove the last part of the statement, we write

    R|gx(y)|qdy=R(1,1)(1|x|m)s(p1)q|xy|(1+sp)qdy+11|gx(y)|qdy.

    The first integral in the sum is finite, being x(1,1) fixed, and y(1,1). Concerning the second one, we re-write it arguing as in the first part of this proof

    11|gx(y)|qdy=11|fx(t)|qdt,

    and use that |fx(t)|qc(x)q|t|(1+sp(p1))q as t0, to conclude that gx(y)Lq(R) whenever (2p(1s))q<1. In particular, gx(y)Lq(R) for every q1, if 2p(1s)<0. We need to show now that the following integral is finite

    RR|gx(y)gx(z)|q|yz|1+rqdydz=R(1,1)R(1,1)dydz+2R(1,1)11dydz=+1111dydz=:I1+2I2+I3

    for some r. To this aim, we observe that the most singular case is when y,z(1,1), and both yx and zx. Therefore, we restrict the study only to the last integral in the sum above:

    I3=x1(11|gx(y)gx(z)|q|yz|1+rqdy)dz+1x(11|gx(y)gx(z)|q|yz|1+rqdy)dz. (4.5)

    We consider the first inner integral in dy. For every z(1,x) fixed, and yx:

    gx(y)sgn(x)(ms|x|m1(1|x|m)1s)p1|yx|p2(yx)|yx|1+sp=:c(x)sgn(yx)|yx|p(1s)2, (4.6)

    and so, being 2p(1s)<0,

    |gx(y)gx(z)|q|yz|1+rq|gx(z)|q|xz|1+rqas yx.

    Thus, integrating now in dz, we have that the integral x1(11|gx(y)gx(z)|q|yz|1+rqdy)dz has the same behavior of

    x1|gx(z)|q|xz|1+rqdz.

    Now, like in (4.6), as zx,

    |gx(z)|q|xz|1+rq|c(x)sgn(zx)|zx|p(1s)2|q|xz|1+rq=|c(x)|q|xz|1+rq+(2p(1s))q.

    Hence, the first double integral in (4.5) is convergent, being 1+rq+(2p(1s))q<1 by assumption. The proof of the convergence of the second integral is similar and we omit it.

    In this Section we show that p>2 and s(0,1) such that (1.2) is not constant in (1,1). To this aim it is sufficient to show that

    I(s,p)(x)=limε0(Bε(x))cgx(y)dy (5.1)

    is not constant, where gx is the function defined in (4.4). We will limit ourselves to provide numerical evidence to this statement.

    For sake of clearness, we omit now the indices s and p in I(s,p)(x), noticing that the approximations we are going to present are valid for any s and p for which I(x)=I(s,p)(x) is finite. Then we split I(x)=I(s,p)(x) into the sum of six contributions as follows:

    I(x)=1gx(y)dyI1(x)+xδ1gx(y)dyI2(x)+xxδgx(y)dyI3(x)+x+δxgx(y)dyI4(x)+1x+δgx(y)dyI5(x)++1gx(y)dyI6(x), (5.2)

    where δ>0 will be specified later.

    The most challenging integrals to compute are I3(x) and I4(x) because of the presence of the singularity of gx(y) at y=x.

    From now on, we denote by ˜Ik(x) the numerical approximation of the integral Ik(x) for k=1,,6.

    The integrals I1(x) and I6(x) are approximated by an adaptive quadrature formula implemented in the functions integral and quadva of MATLAB [21], after operating a change of variable that transforms them to integrals on a finite interval with a very mild singularity. The approximated integrals ˜I1(x) and ˜I6(x) are computed by ensuring that

    |Ik(x)˜Ik(x)|1015 for k{1,6}. (5.3)

    Since we are performing our computations with double-precision arithmetic for which the machine precision is about 1016, the tolerance of 1015 in (5.3) is fully satisfactory.

    The approximate integrals ˜I2(x),,˜I5(x) are computed by the Gauss–Legendre quadrature formula using (n+1) nodes (see, e.g., [2,(2.3.10)]). To highlight the dependence of the computed integrals on the number of nodes, we use the notation ˜Ik,n(x) instead of ˜Ik(x), for k{2,3,4,5}.

    For what concerns the numerical error of the Gauss–Legendre quadrature formula, it is possible to prove that there exists a positive constant C only depending on the size of the integration interval such that, for k=2,,5 and for any x(1,1), it holds

    |Ik(x)˜Ik,n(x)|Cnσ (5.4)

    provided that g_x\in W^{\sigma, 2}(\Lambda_k) for some \sigma > 1/2 and where \Lambda_k denotes the integration interval of the integral I_k(x) . The proof of (5.4) follows by applying the estimate (5.3.4a) of [2] and the estimate (3.7) of [3] with Legendre weight w(y)\equiv 1 .

    Then, thanks to the estimates (5.3) and (5.4), it holds that the global approximated integral

    \begin{equation} \tilde I(x) = \sum\limits_{k = 1}^6 \tilde I_k(x) \end{equation} (5.5)

    satisfies the estimate

    \begin{equation} |I(x)-\tilde I(x)|\leq c n^{-\sigma } \|g_x\|_{W^{\sigma, 2}(\Lambda_k)}+10^{-15}, \end{equation} (5.6)

    i.e., \tilde I(x) converges to the exact value I(x) when n\to \infty , for any x\in(-1, 1) up to the tolerance \overline\epsilon = 10^{-15} .

    To get it, it is sufficient to take a number (n+1) of quadrature nodes sufficiently large to guarantee that the error |I(x)-\tilde I(x)| be small enough. Since the value of I(x) is unknown when p\neq 2 , but it is known when p = 2 , we take the case p = 2 as a playground to learn how many quadrature nodes we need to consider in order to approximate I(x) with the desired accuracy.

    Let us now resume the original notation of I^{(s, p)}(x) because we are interested in distinguishing what happens for different values of p and s .

    All the numerical results that will be reported in the next sections were obtained using the MATLAB functions available on the github repository [11].

    When p = 2 we know that (see the proof of Theorem 3.2)

    \begin{equation} I^{(s, 2)}(x) = \frac{\pi}{\sin(\pi s)}. \end{equation} (5.7)

    In Figure 1, left, we plot the values \tilde I^{(s, 2)}(x) , for several values of x\in(-1, 1) and for s\in\{0.2, \ 0.4, \ 0.5, \ 0.58\overline{3}\} . We have chosen \delta = 1/50 in (5.2). Numerical results are fully consistent with the theoretical result reported in (5.7), the values \frac{\pi}{\sin(\pi s)} are represented by the empty squares (only in correspondence of x = 0 ).

    Figure 1.  On the left, the approximated integrals \tilde I^{(s, 2)}(x) , the empty squares at x = 0 represent the values (5.7). On the right, the absolute errors |I^{(s, 2)}(x)-\tilde I^{(s, 2)}(x)| .

    In Figure 1, right, we report the absolute errors |I^{(s, 2)}(x)-\tilde I^{(s, 2)}(x)| for several values of x\in(-1, 1) . When s = 0.2 , s = 0.4 , and s = 0.5 , the errors are all below 5\cdot 10^{-7} . Instead, when s = 0.58\overline{3} , the errors are about 10^{-6} in the middle of the interval and reach the value 10^{-4} when |x| tends to 1. We explain this behavior to the fact that when s\to 1^- , the order of infinity of the function g_x(y) at y = x increases and the computation of the corresponding integral is very demanding.

    In Figure 2 we show the behavior of the errors |I^{(s, 2)}(x)-\tilde I^{(s, 2)}(x)| versus n , and for different values of s , at x = 0 (left) and x = 0.5 (right).

    Figure 2.  The absolute errors |I^{(s, 2)}(x)-\tilde I^{(s, 2)}(x)| versus n for different values of s . On the left at x = 0 , on the right at x = 0.5 .

    When p = 2 there is no value of s > 1/2 for which we know that g_x\in W^{s, 2}({\mathbb R}) (see Proposition 4.3), hence we cannot take advantage of the estimate (5.4). Yet, we observe that the errors for all the values of s decrease when n grows up, showing convergence of the approximated integrals to the exact ones. The value n = 256 provides very satisfactory results: all the errors are lower than 10^{-6} . Moreover, we can conclude that the accuracy of the quadrature formula at x = 0 and x = 0.5 is almost the same for n ranging between 64 and 256 .

    So far, we have tested the accuracy of our quadrature formulas; now we can move to the case p\neq 2 , for which we only know the exact value of the integral I^{(s, p)}(x) when x = 0 . As a matter of fact, we have (see (4.3))

    \begin{equation} I^{(s, p)}(0) = \frac{2}{sp}+2\int_0^1\frac{(1-(1-y^m)^s)^{p-1}}{y^{1+sp}}dy \end{equation} (5.8)

    and we have computed it symbolically by Wolfram Mathematica [22].

    In Figure 3, left, we report the values of I^{(s, p)}(x) when p = 3 , for five values of s and different values of x\in(-1, 1) . Clearly, I^{(s, p)}(x) is not constant in (-1, 1) . The square symbols at x = 0 represent the exact values (5.8). In the right picture of Figure 3 we display the errors |I^{(s, 3)}(0)-\tilde{I}^{(s, 3)}(0)| for five values of s versus the parameter n (related to the number of quadrature nodes). When n increases all the errors decrease with rate comparable with that for the case p = 2 (see Figure 2). Then we expect that the same accuracy occurs in correspondence to other points x\neq 0 that stand sufficiently far from the end-points of the interval (-1, 1) . Differently than for the case p = 2 , here we have reported numerical results also for s = 2/15 , so that g_x\in W^{r, 2}({\mathbb R}) with r = 0.6 , and for which the estimate (5.4) holds.

    Figure 3.  On the left, the approximated integrals \tilde I^{(s, 3)}(x) . The empty squares at x = 0 represent the exact values (5.8). On the right, the absolute errors |I^{(s, 3)}(x)-\tilde I^{(s, 3)}(x)| at x = 0 . The numerical integrals are evaluated using (n+1) nodes.

    Similar results, but now for p = 4 , are shown in Figure 4: on the left, we report the values of I^{(s, 4)}(x) for four values of s and different values of x\in(-1, 1) . Also in this case it is evident that I^{(s, 4)}(x) is not constant in (-1, 1) . The square symbols at x = 0 refer to the exact values (5.8). In the right picture of Figure 4 we show the errors |I^{(s, 4)}(0)-\tilde{I}^{(s, 4)}(0)| for four values of s versus the parameter n (related to the number of quadrature nodes). Similar conclusions made for p = 3 can be drawn for p = 4 , too.

    Figure 4.  On the left, the approximated integrals \tilde I^{(s, 4)}(x) . The empty squares at x = 0 represent the exact values (5.8). On the right, the absolute errors |I^{(s, 4)}(x)-\tilde I^{(s, 4)}(x)| at x = 0 . The numerical integrals are evaluated using (n+1) nodes.

    Bearing in mind that when p = 2 the errors at x = 0 and x = 0.5 were substantially the same, for a fixed value of s , we can conclude that also when p > 2 the accuracy in approximating the integrals at x\neq 0 is comparable to that obtained at x = 0 . Moreover, we observe that, for a fixed s , the regularity of g_x(y) increases with p and this allows us to benefit of the greater convergence order in the estimate (5.4). This implies that, when p > 2 , we can expect that the approximated integrals are at least accurate as those for p = 2 .

    In conclusion, in Table 1 we report the approximated values \tilde I^{(s, p)}(x) for p = 3 and p = 4 , for some values of s and at the two points x = 0 and x = 0.5 . Because these values approximate the corresponding exact values with errors lower than about 10^{-6} , we can state once more that \exists\, p\neq 2 and \exists \, s\in(0, 1) such that I^{(s, p)} is not constant in (-1, 1) .

    Table 1.  The values of \tilde I^{(s, p)}(0) and \tilde I^{(s, p)}(0.5) for some values of s , computed with the formula (5.5) and n = 256 . On the left p = 3 , on the right p = 4 . These values approximate the corresponding exact values with errors lower than 5\cdot 10^{-5} .
    s \tilde I^{(s, 3)}(0) \tilde I^{(s, 3)}(0.5) s \tilde I^{(s, 4)}(0) \tilde I^{(s, 4)}(0.5)
    0.1\overline{3} 5.0446 4.8644 0.1\overline{3} 3.7625 3.4608
    0.20 3.4253 3.2945 0.20 2.5335 2.3025
    0.40 1.9911 2.0046 0.40 1.4166 1.3743
    0.50 1.8484 1.9451 0.50 1.2876 1.3469
    0.58\overline{3} 1.8891 2.0702 0.58\overline{3} 1.2962 1.4584

     | Show Table
    DownLoad: CSV

    F. C. was partially supported by the GNAMPA – INdAM Project 2020 "Problemi ai limiti per l'equazione della curvatura media prescritta". F. F. is a member of GNAMPA – INdAM. P. G. was partially supported by GNCS – INdAM.

    The authors declare no conflict of interest.



    [1] X. Cabré, Y. Sire, Nonlinear equations for fractional Laplacians, Ⅰ: Regularity, maximum principles, and Hamiltonian estimates, Ann. Inst. H. Poincaré Anal. Non Linéaire, 31 (2014), 23–53. http://dx.doi.org/10.1016/j.anihpc.2013.02.001 doi: 10.1016/j.anihpc.2013.02.001
    [2] C. Canuto, M. Y. Hussaini, A. Quarteroni, T. A. Zang, Spectral methods: Evolution to complex geometries and applications to fluid dynamics, Berlin, Heidelberg: Springer, 2007. http://dx.doi.org/10.1007/978-3-540-30728-0
    [3] C. Canuto, A. Quarteroni, Approximation results for orthogonal polynomials in Sobolev spaces, Math. Comput., 38 (1982), 67–86. http://dx.doi.org/10.1090/S0025-5718-1982-0637287-3 doi: 10.1090/S0025-5718-1982-0637287-3
    [4] W. Chen, C. Li, Maximum principles for the fractional p-Laplacian and symmetry of solutions, Adv. Math., 335 (2018), 735–758. http://dx.doi.org/10.1016/j.aim.2018.07.016 doi: 10.1016/j.aim.2018.07.016
    [5] E. Cinti, F. Colasuonno, A nonlocal supercritical Neumann problem, J. Differ. Equations, 268 (2020), 2246–2279. http://dx.doi.org/10.1016/j.jde.2019.09.014 doi: 10.1016/j.jde.2019.09.014
    [6] F. Colasuonno, F. Ferrari, The soap bubble-theorem and a p-Laplacian overdetermined problem, Commun. Pure Appl. Anal., 19 (2020), 983–1000. http://dx.doi.org/10.3934/cpaa.2020045 doi: 10.3934/cpaa.2020045
    [7] L. M. Del Pezzo, A. Quaas, A Hopf's lemma and a strong minimum principle for the fractional p-Laplacian, J. Differ. Equations, 263 (2017), 765–778. http://dx.doi.org/10.1016/j.jde.2017.02.051 doi: 10.1016/j.jde.2017.02.051
    [8] F. del Teso, D. Gómez-Castro, J. L. Vázquez, Three representations of the fractional p-Laplacian: semigroup, extension and Balakrishnan formulas, Fract. Calc. Appl. Anal., 24 (2021), 966–1002. http://dx.doi.org/10.1515/fca-2021-0042 doi: 10.1515/fca-2021-0042
    [9] B. Dyda, Fractional calculus for power functions and eigenvalues of the fractional Laplacian, Fract. Calc. Appl. Anal., 15 (2012), 536–555. http://dx.doi.org/10.2478/s13540-012-0038-8 doi: 10.2478/s13540-012-0038-8
    [10] M. M. Fall, S. Jarohs, Overdetermined problems with fractional Laplacian, ESAIM Control Optim. Calc. Var., 21 (2015), 924–938. http://dx.doi.org/10.1051/cocv/2014048 doi: 10.1051/cocv/2014048
    [11] P. Gervasio, A MATLAB tool for evaluating the fractional p-Laplace operator in 1D domains. Available from: https://github.com/pgerva/fractional-p-laplace.git.
    [12] A. Iannizzotto, S. Mosconi, N. Papageorgiou, On the logistic equation for the fractional p-laplacian, 2021, arXiv: 2101.05535.
    [13] A. Iannizzotto, S. Mosconi, M. Squassina, Global Hölder regularity for the fractional p-Laplacian, Rev. Mat. Iberoam., 32 (2016), 1353–1392.
    [14] H. Ishii, G. Nakamura, A class of integral equations and approximation of p-Laplace equations, Calc. Var., 37 (2010), 485–522. http://dx.doi.org/10.1007/s00526-009-0274-x doi: 10.1007/s00526-009-0274-x
    [15] S. Jarohs, Strong comparison principle for the fractional p-Laplacian and applications to starshaped rings, Adv. Nonlinear Stud., 18 (2018), 691–704. http://dx.doi.org/10.1515/ans-2017-6039 doi: 10.1515/ans-2017-6039
    [16] S. Jarohs, T. Weth, On the strong maximum principle for nonlocal operators, Math. Z., 293 (2019), 81–111. http://dx.doi.org/10.1007/s00209-018-2193-z doi: 10.1007/s00209-018-2193-z
    [17] E. Lindgren, P. Lindqvist, Fractional eigenvalues, Calc. Var., 49 (2014), 795–826. http://dx.doi.org/10.1007/s00526-013-0600-1 doi: 10.1007/s00526-013-0600-1
    [18] J. J. Manfredi, p-harmonic functions in the plane, Proc. Amer. Math. Soc., 103 (1988), 473–479. http://dx.doi.org/10.1090/S0002-9939-1988-0943069-2 doi: 10.1090/S0002-9939-1988-0943069-2
    [19] R. Musina, A. I. Nazarov, Strong maximum principles for fractional Laplacians, Proc. Roy. Soc. Edinb. A, 149 (2019), 1223–1240. http://dx.doi.org/10.1017/prm.2018.81 doi: 10.1017/prm.2018.81
    [20] S. G. Samko, A. A. Kilbas, O. I. Marichev, Fractional integrals and derivatives: theory and applications, Yverdon: Gordon and Breach Science Publishers, 1993.
    [21] L. F. Shampine, Vectorized adaptive quadrature in Matlab, J. Comput. Appl. Math., 211 (2008), 131–140. http://dx.doi.org/10.1016/j.cam.2006.11.021 doi: 10.1016/j.cam.2006.11.021
    [22] Wolfram Research Inc., Mathematica, Version 12.3.1, Champaign, IL, 2021. Availabe from: https://www.wolfram.com/mathematica.
  • This article has been cited by:

    1. Anderson L. A. de Araujo, Edir J. F. Leite, Aldo H. S. Medeiros, Principal curves to fractional m-Laplacian systems and related maximum and comparison principles, 2024, 27, 1311-0454, 1948, 10.1007/s13540-024-00293-1
  • Reader Comments
  • © 2023 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(2355) PDF downloads(176) Cited by(1)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog